Вы находитесь на странице: 1из 7

CFD analysis of the HyShot II scramjet combustor

C. Fureby
a,
, M. Chapuis
a
, E. Fedina
a
, S. Karl
b
a
Defense Security Systems Technology, The Swedish Defense Research Agency FOI,
SE 147 25 Tumba, Stockholm, Sweden
b
Institute of Aerodynamics and Flow Technology, German Aerospace Center, DLR, 37073 Goettingen, Germany
Available online 21 September 2010
Abstract
The development of novel air-breathing engines such as supersonic combustion ramjets (scramjets)
depends on the understanding of supersonic mixing, self-ignition and combustion. These aerothermochem-
ical processes occur together in a scramjet engine and are notoriously dicult to understand. In the present
study, we aim at analyzing the HyShot II scramjet combustor mounted in the High Enthalpy Shock Tunnel
Go ttingen (HEG) by using Reynolds Averaged Navier Stokes (RANS) and Large Eddy Simulation (LES)
models with detailed and reduced chemistry. To account for the complicated owin the HEGfacility a zonal
approach is adopted in which RANS is used to simulate the owin the HEGnozzle and test-section, provid-
ing the necessary inow boundary conditions for more detailed RANS and LES of the reacting ow in the
HyShot combustor. Comparison of predicted wall pressures and heat uxes with experimental data show
good agreement, and in particular does the LES agree well with the experimental data. The LES results
are used to elucidate the ow, mixing, self-ignition and subsequent combustion processes in the combustor.
The combustor ow can be separated into the mixing zone, in which turbulent mixing from the jet-in-cross
ow injectors dominates, the self-ignition zone, in which self-ignition rapidly takes place, and the turbulent
combustion zone, located towards the end of the combustor, in which most of the heat release and volumetric
expansion takes place. Self-ignition occurs at some distance downstream of the injectors, resulting in a dis-
tinct pressure rise further downstream due to the volumetric expansion as observed in the experiments.
The jet penetration is about 30%of the combustor height and the combustion eciency is found to be around
83%.
2010 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
Keywords: Scramjet; HyShot II wind-tunnel experiments; Large Eddy Simulation; Supersonic mixing; Self-ignition
1. Introduction
The development of reliable hypersonic ight
vehicles requires the solution of many technical
challenges associated with the comparatively small
net thrust at supersonic or hypersonic ight speeds.
One of the more essential issues is the design of an
air-breathing propulsion systemcapable of operat-
ing over the wide range of Mach (Ma) numbers
desired to facilitate the advancement of high-speed
ight and space access vehicles. For ights in the
supersonic (3 < Ma < 5) regime a ramjet, in which
the ow is decelerated to subsonic levels before it
enters the combustor, is favored whereas for ights
in the hypersonic (5 < Ma < 15) regime a super-
sonic combustion ramjet (scramjet), in which ow
through the engine remains supersonic, is pre-
ferred, [1]. Below Ma 3 a turbojet engine can be
1540-7489/$ - see front matter 2010 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.proci.2010.07.055

Corresponding author. Fax: +46 8 5550 4144.


E-mail address: fureby@foi.se (C. Fureby).
Available online at www.sciencedirect.com
Proceedings of the Combustion Institute 33 (2011) 23992405
www.elsevier.com/locate/proci
Proceedings
of the
Combustion
Institute
used to accelerate the vehicle to ramjet speed,
whereas a rocket motor is required to accelerate
the vehicle to scramjet speed. To facilitate the con-
tinuous operation from subsonic to hypersonic
speeds dual-mode ramjet/scramjet engines have
been proposed [2]. Although the ramjet technology
is mature enough for working engines, the scramjet
technology, and in particular the scramjet combus-
tor technology [3], is critically relaying onimproved
understanding of the aerothermochemistry in the
scramjet engine.
At scramjet conditions fuel is supplied to the
supersonic air-stream through the combustor
whereafter mixing, self-ignition and combustion
are supposed to complete the process of super-
sonic combustion. Most experimental research
on supersonic combustion has been conducted in
ground based research facilities [46], but recently
a few scramjet powered vehicles have own for a
few seconds [78], generating pioneering data. In
this study, we use numerical simulations to repro-
duce and analyze, in detail, the processes taking
place in the HyShot II combustor under ight-like
conditions. Flight test data [9], and experimental
data from the High Enthalpy Shock Tunnel Go t-
tingen (HEG) of the German Aerospace Centre
(DLR) [10], in particular, will aid in validating
the computational models and aid in the analysis.
Reynolds Averaged Navier Stokes (RANS) mod-
els, using presumed-shape Probability Density
Function (PDF) combustion models [11], and
Large Eddy Simulation (LES) models, using the
Partially Stirred Reactor (PaSR) turbulence chem-
istry interaction model [12], will complement each
other in simulating the ow in the HEG nozzle
and test-section, and in the HyShot II scramjet
combustor. Accurate predictions of mixing, self-
ignition and subsequent combustion in the
HyShot II scramjet combustor requires that the
intake ow is accurately predicted, and thus that
both the nozzle and test sections are included in
the numerical model. A zonal approach is
adopted in which RANS is used to predict the
ow in the nozzle and test sections thus providing
accurate inow boundary conditions to the more
detailed combustor RANS and LES.
2. HyShot HEG experimental studies and ight
tests
In the HyShot II ight experiment [8], a two-
stage Terrier-Orion Mk70 sounding rocket carried
the payload to an apogee of 315 kmafter which the
vehicle was turned and gravitationally accelerated
during the descent phase. Hydrogen (H
2
) super-
sonic combustion data was taken at ight Ma num-
bers between M = 7.6 and 7.8 in an altitude range
from 35 km down to 23 km. Post-ight analysis
was performed in the free piston shock tunnel T4
at the University of Queensland [9], and in the
HEG [10], and here we consider HEG test condi-
tion XII, representing the HyShot II freestream
conditions at an altitude of 32.5 km. The owpath
geometry of the wind-tunnel model is identical to
the ight conguration and consists of an intake
and two combustion chambers as shown in Fig. 1.
The intake ramp is a wedge of 18 half angle and
is 100 mm wide whereas the cross-sectional dimen-
sions of the combustion chambers are
9.8 75 mm
2
. Gaseous hydrogen is injected per-
pendicular to the combustor air-streamat four fuel
injectors 58 mm downstream of the combustion
chamber leading edge. Each combustion chamber
is 232 mm long, and attached to a single-sided
expansion forming the thrust surface. Between
the intake ramp and the combustion chamber, side
and oor bleeds are used to spill the boundary lay-
ers from the oor and the side-walls of the intake,
allowing the leading edge cowl shocks to pass the
combustion chamber. Compared to the ight tests
the HEG experiments oers the possibility of con-
trolled conditions and more detailed measure-
ments, which is the main reason for selecting the
HyShot experiments as the target for this study.
In the HEG combustor, 14 pressure transducers
and 10 thermocouples (cf. Fig. 1) are used to mea-
sure the pressure and wall heat-ux.
3. Mathematical and numerical modeling
Due to the complexity of simulating the react-
ing ow in the HyShot scramjet in the HEG facil-
Fig. 1. The HyShot II wind-tunnel model with fuel injectors and data acquisition system.
2400 C. Fureby et al. / Proceedings of the Combustion Institute 33 (2011) 23992405
ity under ight-like conditions, a zonal approach
has been adopted [13], in which dierent simula-
tions have been performed for the dierent parts
of the HEG facility as shown in Fig. 2a. The rst
simulation aims at evaluating the HEG nozzle
ow conditions (velocity, pressure, temperature
and composition) discharging into the HEG test-
section, where the HyShot model is mounted. This
simulation was performed in an axisymmetric
domain (with 20,000 cells) using the non-equilib-
rium ve-species, ve-step reaction mechanism
of Gupta et al. [14], and a steady-state compress-
ible RANS model implemented in the TAU code
[15]. The results agree well with experimental data
and provide the boundary conditions for the sec-
ond simulation of the HEG test-section contain-
ing the HyShot wind-tunnel model [13], which
was performed in the planar domain of Fig. 2a,
with 115,000 cells, adapted to the resulting shock
system and boundary layers. The results from the
rst (nozzle ow) simulation are used as fareld
boundary conditions for the second (test-section)
simulation in which the 18 inclined intake ramp
and combustion chamber are modeled as no-slip
and isothermal walls. Wave-transmissive outow
boundary conditions [16], are used at the oor
bleed outlet and at the HyShot scramjet combus-
tor outlet. To emulate ight conditions the entire
conguration was mounted at 3.8 angle-of-
attack. The same computational model as used
for the rst (nozzle ow) simulation is used. The
velocity, temperature, pressure and composition
proles at the cut-plane x = 0.360 m, shown in
Fig. 2b, will act as the inow conditions to the
HyShot combustor, the analysis of which is the
primary objective of this study. For the HyShot
combustor simulations both RANS and LES
models have been used to capture dierent aspects
of the physics and to evaluate these models.
The RANS and LES models are based on the
same reacting ow model in which the mixture is
assumed to be a linear viscous uid with Fourier
heat conduction and Fickian species diusion. In
the RANS Blottner curve-ts with the Wilke mix-
ture rule and a modied Eucken correction with
the HerningZipperer mixture rule were applied
for the calculation of viscosity and thermal conduc-
tivity [13]. In the LES, the viscosity is computed
according to Sutherlands lawand the thermal con-
ductivity and species diusivities are computed
using the viscosity and constant Prandtl and species
Schmidt numbers, respectively [16]. The mixture
thermal and caloric equations of state are obtained
under the assumption that each species is a ther-
mally perfect gas, with tabulated formation enthal-
pies and specic heats, respectively. The reaction
rates are computed from GuldbergWaages law
of mass action by summation over all participating
reactions, with rate constants obtained from a
modied Arrhenius law. For RANS, the nine-spe-
cies, 19-step Jachimowski mechanism [17], is used
whereas for the LES the seven-species, eight-step
mechanism of Davidenko et al. mechanism [18], is
used, resulting in the same ignition delay time
prediction.
The RANS model used [11], is based on the
ensemble averaged version of the governing equa-
tions with the Reynolds stress tensor and ux vec-
tors modeled by the two-equation Wilcox kx
and the one-equation SpalartAllmaras models
[19]. Here, only the SpalartAllmaras results are
discussed. The ensemble averaged reaction rates,
including the turbulence chemistry interaction
terms, are modeled by a presumed-shape PDF
model based on a Gaussian distribution for the
temperature and a multivariate b-PDF for the spe-
cies mass fractions [11]. In addition to the ensemble
averaged mass, momentum, energy, species and
turbulence equations this requires the solution of
two additional equations for the variances of the
temperature and the sum of the species mass frac-
tions [20]. The equations are solved using an expli-
cit multi-stage Runge-Kutta scheme using a point-
implicit treatment of the chemical source terms and
local time stepping for convergence acceleration.
The AUSMDV ux vector-splitting scheme and
the MUSCL gradient reconstruction algorithms
are employed.
Fig. 2. Computational models. (a) Schematic of the HEG and zonal partitioning of the computational model and (b) the
resulting combustor inow velocity, pressure and temperature proles. The proles are normalized by v
0
= 1800 m/s,
T
0
= 1459 K and p = 53.0 kPa.
C. Fureby et al. / Proceedings of the Combustion Institute 33 (2011) 23992405 2401
The LES model used [12], is based on the low-
pass ltered version of the governing equations
with the subgrid stress tensor and ux vectors
modeled by the mixed model [21]. The ltered
reaction rates, including the subgrid turbulence
chemistry interaction terms, are here modeled by
the PaSR model [22], in which the ltered reaction
rates are evaluated as the product of the reaction
rates of the ne structures and their volume frac-
tion. This model has also been carefully validated,
cf. [12] and references therein, with very good
results for a wide range of applications. The equa-
tions are solved using a fully explicit nite volume
discretization, based on the C++ library open-
Foam [23], utilizing a monotonicity preserving
ux reconstruction algorithm [24], and two-stage
Runge-Kutta time integration.
4. The HyShot II supersonic combustor model
The computational model of the HyShot scram-
jet combustor used is presented in Fig. 3. At the
inow (x = 0.360 m) Dirichlet conditions are used
for all variables with proles (Fig. 2a) resulting
fromthe test-section simulation. Isothermal no-slip
wall boundary conditions are used to represent the
top and bottom walls of the combustor whereas
symmetry conditions are used in the spanwise
direction, thus neglecting the inuence of the side-
walls. The wall temperature is xed to 300 K to
account for the short test time in the HEG facility.
The exhaust nozzle was only partially included up
to a plane 0.110 m downstream of the end of the
combustion chamber (at x = 0.760 m) at which all
variables are extrapolated. Fuel injection was mod-
eled by partially including the injectors using a total
hydrogen pressure of 2.97 bar and a total tempera-
ture of 300 K at the injector inow. For the HEG
experiment simulations the global equivalence
ratio was / 0:43 but for the ight tests it was
slightly lower. The RANS computations make
use of all symmetries of the combustor and hence
only one quarter of the computational domain
shown in Fig. 3 was modeled, whereas in the LES
the domain shown in Fig. 3 is modeled. For RANS,
unstructured grids, adapted to the density, temper-
ature and pressure gradients in the ow, are used
with about 0.8 million points, whereas for LES a
coarse and a ne grid with 12.5 and 26.2 million
points, respectively, are used. For both RANS
and LES performed grid renement studies indi-
cate that the grids used are suciently ne to cap-
ture the key ow features.
5. Results from combustor simulations
The composite Fig. 4 presents selected aspects
of the reacting ow in the HyShot II combustor
from the LES computations. The wall pressure in
Fig. 4a reveals that initially the pressure increases
slowly with increasing distance fromthe transverse
fuel jets to increase more rapidly between 30 and
60 Ddownstreamof the transverse fuel jets to peak
at about 100 D. Higher wall pressures are also
observed beneath the bow-shock, forming a hood
over the transverse fuel jet and beneath the fuel jets.
The more rapid pressure increase further down-
stream is caused by volumetric expansion due to
exothermicity and chemical reactions. The large
diusivity of H
2
allows it to rapidly mix with the
freestream air, forming a reactive mixture around
the high momentum H
2
lled jet core seen in
Fig. 4c.
The velocity in Fig. 4b shows the high-speed
(ight Ma 7.2) ow entering the combustor,
the transverse H
2
jet, the bow-shock and the com-
plex ow structures developing downstream of the
injection points. The jet-to-freestream momentum
ux ratio, J q
H2
v
2
H2
=q
air
v
2
air
, is 1.1 and the
mean H
2
jet penetration prole is close to that dis-
cussed in [25], and the H
2
jets are observed to pen-
etrate to 30% of the combustor height before
self-ignition (Fig. 5). The axial velocity in the
combustor is somewhat reduced compared with
that at the inow due to the combined blocking
and redirection eects caused by the transverse
jets, helping to stabilize the ame.
The time-averaged transverse H
2
jets typically
consist of a counter-rotating vortex pair and a
horseshoe-vortex, whereas instantaneously they
consist of smaller and topologically more complex
vortex structures as seen in the second invariant, k
2
,
of the velocity gradient in Fig. 4c. These typically
consist of small bent S-shaped vortices (side arms)
with their lower parts aligned with the ow and
their upper parts curling over the jet forming the
Fig. 3. Computational model of the HyShot II combustor used in the RANS and LES analysis.
2402 C. Fureby et al. / Proceedings of the Combustion Institute 33 (2011) 23992405
neck (circumferential rollers) of the counter-rotat-
ing vortex pair. This feature is observed experimen-
tally by Ben-Yakar et al. [25], and they propose that
the side arms are stretched by increased shear stres-
ses in the regions of steep velocity gradient. These
vortical structures seem to arise from Kelvin
Helmholz (KH) instabilities in the jet shear layers
just beneath the bow-shock. Since the transverse
jets contain all the H
2
, mixing dominates during
the rst 20 to 40 D, whereby air is entrained into
the vortex structures and H
2
diuses into the air,
resulting in a combustible mixture around the jets.
Further downstream, between 30 and 60 D the H
2
and air are suciently mixed to burn if the temper-
ature is suciently high. In this region, self-ignition
occurs intermittently in lean low vorticity regions
with temperatures close to the laminar self-ignition
temperature but also due to shockshock interac-
tions occurring at lower temperatures. The volu-
metric expansion causes the S-shaped side arms
and spanwise rollers to combine into X-shaped vor-
tices, dominating the self-ignition region. Due to
volumetric expansion, vortex stretching, baroclinic
torque and self-diusion, the vortex structures
eventually develop into longitudinal vortices, dom-
inating the downstream part of the combustor.
These vortices growin size with increasing distance
from the injection point due to the volumetric
expansion, and when they reach the end of the com-
bustor, the gradual expansion increases the veloc-
ity, presented in Fig. 4b, causing a forward
directed thrust on the thrust surface.
In Fig. 4d, the iso-surfaces of the H
2
mass frac-
tion (gray) and the heat release, dened as the
source term in the transport equation for the sen-
sible enthalpy [26], conditioned on k
2
supports the
previous description of the self-ignition process.
Heat release occurs locally beneath the bow-shock
as H
2
and air are mixed when the transverse H
2
jets impinge on the airow through the combus-
tor. However, the heat release ceases further
downstream, between 10 and 25 D, due to insu-
cient mixing. Further downstream, where H
2
and
air are well mixed, self-ignition, aided by recircu-
Fig. 4. Composite gure of the reacting ow in the HyShot II combustor: (a) wall pressure and an iso-surface of the H
2
mass fraction, (b) axial velocity cut through a fuel injector, (c) iso-surface of the second invariant of the velocity gradient,
k
2,
, colored by the temperature and (d) iso-surfaces of the H
2
mass fraction (gray) and the heat release conditioned on k
2
.
Fig. 5. Comparison of LES and RANS predictions in terms of contours of the instantaneous and mean axial velocity, ~v
x
and hv
x
i, and temperature,
~
T and hTi, and numerical schlieren images superimposed on semi-transparent H
2
mass
fraction distributions. Shown is also a frequency spectrum of the near-wall pressure at three locations (P1: x = 0.45 m,
P2: x = 0.55 m and P3: x = 0.65 m).
C. Fureby et al. / Proceedings of the Combustion Institute 33 (2011) 23992405 2403
lated combustion products, due mainly to inter-
acting shocks, occurs, causing the S-shaped vorti-
ces and the spanwise rollers to rapidly develop X-
shaped vortices. The chain-branching step
H + O
2
MOH + O is particularly important for
self-ignition as its products react with H
2
to pro-
duce H, that continues to react, producing more
radicals, until the resulting pool of radicals
reaches a critical level, whereby a rapid exother-
mic reaction occurs.
In Fig. 5, results from LES and RANS predic-
tions are compared. The perspectives show con-
tours of the instantaneous velocity, ~v
x
, (rear) and
temperature,
~
T, (front) from LES (top) and the
mean axial velocity, hv
x
i, (rear) and temperature,
hTi, (front) from RANS (bottom), at 11 cross-sec-
tions. The time-averaged LES result in similar con-
tours as obtained by RANS and is therefore not
presented. From the LES results the unsteady nat-
ure of the owis evident, with the jets resulting in a
complex wake that interacts with the surrounding
ow. Onaverage, eachjet results ina velocity decit
extending throughout the combustor in RANS
whereas in LES it dissolves at about three quarters
of the combustor. The cold H
2
jets extend further
aft in the RANS results compared to the LES
results in which they break-up and self-ignite ear-
lier. In both cases a thin boundary layer can be seen
at both the top and bottom walls. The side views
show numerical schlieren images, obtained from
the vertical gradient of the refraction index com-
puted from the LorenzLorenz equation [27],
superimposed with semi-transparent contours of
the H
2
mass fraction. From the schlieren image
the bow-shock is clearly visible as is the Ma disc
barrel shock and separated boundary layer just in
front of the transverse H
2
jet. The boundary layers
show signs of intermittent separation, and the jet
penetration is larger in the LES predictions com-
pared to the RANS predictions. The transverse
H
2
jet seems more diluted fromthe LES predictions
than from the RANS predictions but is also
unsteady and strongly three-dimensional. The fre-
quency spectra of the near-wall pressure at P1
(x = 0.45 m), P2 (x = 0.55 m) and P3
(x = 0.65 m) along the jet centerline are also shown
in Fig. 5. Close to the injectors, at P1, the spectrum
consists of several peaks (including the Helmholz
frequency of 12 kHz) whereas further downstream,
at P2, frequencies at 7 kHz (corresponding the to
combustor eigenfrequency as obtained from a
stand-alone combustor eigenmodel calculation)
dominate. Close to the combustor exit (at
x = 0.650 m) the dominating frequency decreases
to about 125 Hz.
In Fig. 6, predicted and measured axial proles
of (a) the time-averaged wall pressure, hpi, and (b)
wall heat-ux, h

hi, are compared. The wall pressure


and heat-ux are compared at lines on the bottom
wall of the combustor between injectors and
6.0 mm o centerline between injectors, respec-
tively. For the time-averaged wall pressure the
experimental data show a sudden increase between
x = 0.50 m and 0.53 m followed by a slower
increase up to the end of the combustor, at
x = 0.65 m, after which the pressure drops rapidly
to the exhaust pressure. The RANS predictions
showan almost linear increase fromthe combustor
inlet to the combustor exit, missing the sharp pres-
sure increase indicating combustion, and under- or
overpredict the wall pressure by up to 25%. The
LES predictions are in better qualitative agreement
with the experimental data, in particular showing a
pressure rise although in two phases and starting
somewhat too early. For the time-averaged wall
heat-ux a sudden rise is observed at about
x = 0.52 m, corresponding well to the location at
which the pressure rises due to combustion. The
average level in the rst part of the combustor
agrees well with the laminar heat-ux predicted
using a Blasius prole and a wall temperature of
300 K, but the transverse H
2
jet introduces some
peculiarities after x = 0.42 m. The RANS predic-
tions typically overpredict the heat-ux whereas
the LES predictions typically underpredict the
heat-ux, with between 15%and 10%, respectively.
Fig. 6. Comparison of predicted and measured (a) wall pressure between jet injectors and (b) heat-ux along a line
6.0 mm o centerline between jet injectors.
2404 C. Fureby et al. / Proceedings of the Combustion Institute 33 (2011) 23992405
Both the pressure and heat-ux predictions are
generally within the experimental uncertainty. In
addition, the combustion eciency, g
c
1
R
A
qj~vj
~
Y
H2
dA= qj~vj
~
Y
H2
A
inj
, at the outlet is found to
be about 83%.
6. Summary and concluding remarks
Here we use RANS and LES to analyze the
HyShot II scramjet combustor at freestream con-
ditions representative of the wind-tunnel experi-
ments in the HEG shock tunnel corresponding
to ight conditions at 32.5 km altitude. To pro-
vide accurate inow boundary conditions to the
reacting RANS and LES combustor simulations
separate RANS computations of the HEG nozzle
and HEG test sections were carried out in
sequence before the reacting RANS and LES
computations were performed. The LES results
reveal a very complicated ow pattern dominated
by the transverse jet-in-cross ow and the associ-
ated fuelair mixing, the self-ignition process
and further aft in the combustor a region with
fully developed turbulent combustion. Between
30 and 60 jet diameters downstream of the injec-
tors self-ignition occurs due to hot spots caused
by colliding shocks supported by recirculated
hot combustion products. This intermittent hot
spot formation leads to a longitudinally oscillat-
ing self-ignition zone that participates in creating
the unsteady features observed in the LES results.
The qualitative average ow features of LES were
found to be similar to the results of the RANS.
Both the qualitative and quantitative agreement
of the surface pressure distribution and surface
heat-ux to the experimental data was improved
by the use of the LES model compared to RANS.
The results show that RANS analysis is sucient
to capture the main ow features at limited com-
putational cost whereas the application of the
LES model can lead to signicant improvement
of the prediction of detailed ow features.
Acknowledgement
The HEG tests were carried out by A. Gardner
and additional funding was provided by the Euro-
pean Space Agency Grant TRP 17001/02/NL/
MV, and the LES computations were supported
by the Swedish Defense Material Agency.
References
[1] E.T. Curran, S.N.B. Murthy (Eds.), Scramjet Pro-
pulsion, Progress in Astronautics and Aeronautics,
vol. 189, AIAA, Washington, DC, USA, 2000.
[2] D. Andreadis, The Industrial Physicist 10 (2007) 26
29.
[3] F. Ladiende (Ed.), AIAA J. 43 (3) (2010), in press
(special Issue on Scramjet Combustion Technology).
[4] W. Waidmann, F. Al, U. Brummund, M. Bo hm,
W. Clauss, M. Oschwald, Space Technol. 15 (1995)
421429.
[5] J.P. Drummond, G.S. Diskin, A.D. Cutler, Fuel
Air Mixing and Combustion in Scramjets, Technol-
ogies For Propelled Hypersonic Flight, Working
Group AVT 10, Final Report, NATO Research and
Technology Organization, 2001.
[6] T. Sunami, P. Magre, A. Bresson, F. Grisch, M.
Orain, M. Kodera, AIAA 2005-3304, 2005.
[7] Available at: <http://www.nasa.gov/missions/
research/x43-main.html>.
[8] A. Paull, H. Alesi, S. Anderson, The HyShot ight
program and how it was developed, in: AIAA
AAAF 11th Int. Space Planes and Hypersonic
Systems and Technologies Conference, Orleans,
France, 2002.
[9] M.K. Smart, N.E. Hass, A. Paull, AIAA J. 44
(2006) 23662375.
[10] J.M. Schramm, S. Karl, K. Hannemann, J. Stree-
lant, AIAA 2008-2547, 2008.
[11] P. Gerlinger, H. Mo bus, D. Bruggemann, J. Com-
put. Phys. 167 (2001) 247276.
[12] C. Fureby, Philos. Trans. R. Soc. A 367 (2009)
29572969.
[13] S. Karl, K. Hannemann, J. Streelant, A. Mack,
AIAA 2006-8041, 2006.
[14] R.N. Gupta, J.M. Yos, R.A. Thompson, K.P. Lee,
A Review of Reaction Rates and Thermodynamic and
Transport Properties for an 11-Species Air Model for
Chemical and Thermal Non-equilibrium Calculations
to 30000 K, NASA RP 1232, 1990.
[15] T. Gerhold, O. Friedrich, J. Evans, M. Galle, AIAA
1997-0167, 1997.
[16] T.J. Poinsot, S.K. Lele, J. Comput. Phys. 101 (1992)
104129.
[17] C.J. Jachimowski, An Analytical Study of the
HydrogenAir Reaction Mechanism with Application
to Scramjet Combustion, NASA TP 2791, 1988.
[18] D.M. Davidenko, I. Go kalp, E. Dufour, P. Magre,
AIAA 2006-7915, 2006.
[19] D.C. Wilcox, Turbulence Modeling for CFD, La
Canada, California, USA, 1998.
[20] S.S. Girimaji, Combust. Sci. Technol. 78 (1991) 177
196.
[21] R. Bensow, C. Fureby, J. Turb. 8 (54) (2007) 117.
[22] M. Berglund, E. Fedina, C. Fureby, V. Sabelnikov,
J. Tegner, Finite rate chemistry les of self ignition in
a supersonic combustion ramjet, AIAA.J., in press,
doi:10.2514/1.43746.
[23] H.G. Weller, G. Tabor, H. Jasak, C. Fureby,
Comput. Phys. 12 (1998) 620631.
[24] D. Drikakis, C. Fureby, F.F. Grinstein, M. Liefen-
dahl, in: F.F. Grinstein, L. Margolin, B. Rider
(Eds.), Implicit Large Eddy Simulation: Computing
Turbulent Fluid Dynamics, Cambridge University
Press, Cambridge, UK, 2007, p. 94.
[25] A. Ben-Yakar, M.G. Mungal, R.K.R.K. Hansen,
Phys. Fluids 18 (2006) 026101.
[26] T. Poinsot, D. Veynante, Theoretical and Numerical
Combustion, Edwards, 2001.
[27] L.A. Vasilev, Schlieren Methods, Keter Inc., New
York, 1971.
C. Fureby et al. / Proceedings of the Combustion Institute 33 (2011) 23992405 2405

Вам также может понравиться